203
37*1 //S/c/ yya, V7fJ COMBINED ELECTROCHEMISTRY AND SPECTROSCOPY OF COMPLEXES AND SUPRAMOLECULES CONTAINING BIPYRIDYL AND OTHER AZABIPHENYL BUILDING BLOCKS DISSERTATION Presented to the Graduate Council of the University of North Texas in Partial Fulfillment of the Requirements For the Degree of DOCTOR OF PHILOSOPHY By Lei Yang, B.S., M.S. Denton, Texas August, 1995

37*1 COMBINED ELECTROCHEMISTRY AND .../67531/metadc279396/...Fig.2.4c Ruthenium(II) complexes which are discussed in Chapter IV 22 Fig.2.5 Platinum(II) and palladium(II) complexes

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

  • 3 7 * 1

    / / S / c /

    y y a , V 7 f J

    COMBINED ELECTROCHEMISTRY AND SPECTROSCOPY OF COMPLEXES

    AND SUPRAMOLECULES CONTAINING BIPYRIDYL AND

    OTHER AZABIPHENYL BUILDING BLOCKS

    DISSERTATION

    Presented to the Graduate Council of the

    University of North Texas in Partial

    Fulfillment of the Requirements

    For the Degree of

    DOCTOR OF PHILOSOPHY

    By

    Lei Yang, B.S., M.S.

    Denton, Texas

    August, 1995

  • 3 7 * 1

    / / S / c /

    y y a , V 7 f J

    COMBINED ELECTROCHEMISTRY AND SPECTROSCOPY OF COMPLEXES

    AND SUPRAMOLECULES CONTAINING BIPYRIDYL AND

    OTHER AZABIPHENYL BUILDING BLOCKS

    DISSERTATION

    Presented to the Graduate Council of the

    University of North Texas in Partial

    Fulfillment of the Requirements

    For the Degree of

    DOCTOR OF PHILOSOPHY

    By

    Lei Yang, B.S., M.S.

    Denton, Texas

    August, 1995

  • 19

    Yang, Lei, Combined Electrochemistry and Spectroscopy of Complexes and

    Supramolecules Containing Bipyridyl and Other Azabiphenyl Building Blocks. Doctor

    of philosophy (Chemistry), August, 1995, 184 pp., 13 tables, 87 illustrations,

    references 196 titles.

    A group of azabiphenyl complexes and supramolecules, and their reduced and

    oxidized forms when possible, were characterized by cyclic voltammetry and

    electronic absorption spectroscopy. The oxidized and reduced species, if sufficiently

    stable, were further generated electrochemically inside a specially designed quartz cell

    with optically transparent electrode, so that the spectra of the electrochemically

    generated species could be taken in situ. Assignments were proposed for both parent

    and product electronic spectra.

    Species investigated included a range of Ru(II) and Pt(II) complexes, as well as

    catenanes and their comparents.

    Using the localized electronic model, the electrochemical reduction can be in

    most cases assigned as azabiphenyl-based, and the oxidation as transition metal-based.

    This is consistent with the fact that the azabiphenyl compounds have a low lying 7t*

    orbital.

    The electronic absorption spectra of the compounds under study are mainly

    composed of n —> K* bands with, in some cases, charge transfer bands also.

  • ACKNOWLEDGMENTS

    It is a privilege to express my sincere appreciation to professor Paul S.

    Braterman, my major advisor. Without his constant instruction, guidance and

    encouragement throughout this work, it would have been impossible to complete all

    the work described here.

    I would very much like to express my gratitude to Dr. Jae-In Song for his

    advice, help and valuable comments. Special thanks also due to Dr. Frank M.

    Wimmer of Universiti Brunei Darussalam, Bandar Seri Begawan, Brunei, for the

    supply of platinum(II) complexes described in Chapter III, and also for his cooperation

    on the work described in that chapter.

    Thanks also due to Drs. G. Brent Young of Imperial College of Science,

    Technology and Medicine, London, UK for the supply of the trimethylsilylmethyl

    complexes and to J. Fraser Stoddart of The University, Sheffield, UK for the supply of

    the catenanes and their precursors.

    I also would like to acknowledge the financial support for this research from

    Robert A. Welch foundation, and the University of North Texas Faculty Research

    Fund.

    m

  • TABLE OF CONTENTS

    PAGE

    LIST OF TABLES viii

    LIST OF ILLUSTRATIONS x

    CHAPTER

    I. INTRODUCTION 1

    1.1 Frontier Molecular Orbitals

    1.2 Supramolecules

    1.3 Azabiphenyl Systems

    1.4 Symmetry Considerations

    H. EXPERIMENTAL 12

    2.1 Materials

    2.2 Electrochemistry and Spectroelectrochemistry

    2.3 The Compounds Under Examination

    2.4 Quantum Chemistry Calculation

    III. ELECTRO- AND SPECTROELECTROCHEMICAL STUDIES OF

    PLATINUM(II) BIPYRIDINE COMPLEXES AND RELATED SPECIES 27

    3.1 Electrochemistry of Bipyridines and Platinum(II) Complexes

    3.1.1 Electrochemistry of Free bipyridines

    3.1.1.1 Electrochemistry of 2,4'-bipyridine

    IV

  • 3.1.1.2 Electrochemistry of N'-Methyl-2,4'-Bipyridine

    3.1.1.3 Electrochemistry of 4,4'-Diphenyl-2,2'-Bipyridine

    3.1.2 Electrochemistry of Platinum(II) Complexes

    3.1.2.1 Ligand-Based Reductions of Platinum(II) Complexes

    [Pt(bipy)ClJ

    [Pt(bipy)(4-NCpy)Cl]+

    [Pt(ph2-bipy)Cy and [Pt(Me2-bipy)Cl2]

    Pt(2,4'-bipyOct-H)Cl2

    and [Pt(2,4'-bipyOct)Cl3]

    [Pt(2,4'-bipyMe-H)(bipy)]2+

    [Pt(2,4'-bipyMe-H)py2]2+

    [Pt(Terpy)Cl]+

    [Pt(DCMB)ClJ

    3.1.2.2 Platinum(II)-Based Reductions

    3.2 Spectroelectrochemistry

    3.2.1 2,4'-Bipyridines

    3.2.2 Spectroelectrochemistry of Square Planar Platinum(II) Complexes with

    2,4'-Bipyridine as Ligand

    [Pt(2,4-bipyOctyl)Cl3]

    [Pt(2,4'-bipyOctyl-H)Cy

    [Pt(bipy)Cy

    [Pt(bipy)(4-CNpy)Cl]+

  • [PtDCMBClJ

    IV. ELECTROCHEMICAL AND SPECTROELECTROCHEMICAL STUDIES OF

    SOME RUTHENIUM(II) AZABIPHENYL COMPLEXES

    AND RELATED SPECIES 75

    4.1 Electrochemical studies

    4.1.1 Electrochemical studies of l,l'-Biisoquinoline

    4.1.2 Electrochemical Studies of 2,2'-Bisquinoline

    4.1.3 Electrochemistry of Ruthenium(II) Complex Ru(bipy)2(biiq)2+

    4.1.4 Electrochemistry of Ruthenium(II) Complexes with Trialkyl Phosphite

    and Trimethylsilylmethyl Ligands

    4.2 Spectroelectrochemistry of [Ru(biiq)(bipy)2]2+ and Related Species

    4.2.1 Spectroelectrochemistry of U'-Biisoquinoline

    4.2.2 Spectroelectrochemistry of [Ru(biiq)(bipy)2]2+

    4.2.3 Spectroelectrochemistry of Ruthenium(II) Complexes with Trialkyl

    Phosphite and Trimethylsilylmethyl Ligands

    V. ELECTROCHEMICAL AND SPECTROELECTROCHEMICAL STUDIES OF

    BIPYRIMIDINE PLATINUM(II) AND PALLADINUM(II) COMPLEXES 108

    5.1 Electrochemistry of Mono- and Dinuclear Platinum(II) and Palladium(II)

    Complexes

    5.1.1 (bipym)Pt(CH2SiMe3)2

    5.1.2 (bipym)Pd(CH2SiMe3)2

    5.1.3 (Me3SiCH2)2Pd(bipym)Pt(CH2SiMe3)2

    VI

  • 5.1.4 (j-bipym(Pd(CH2SiMe3)2)2

    5.2 Spectroelectrochemistry of Mono- and Dinuclear Platinum(II) and

    Palladium(II) Complexes

    VI. ELECTRO- AND SPECTROELECTROCHEMICAL STUDIES OF

    PHENANTHROLINES 127

    6.1 Electrochemistry of Phenanthrolines

    6.2 Spectroelectrochemistry of Phenanthrolines

    Vn. SPECTROELECTROCHEMICAL STUDIES OF SOME PARAQUAT

    CATENANES AND THEIR PRECURSORS 145

    7.1 Electrochemistry of Catenanes and Their Precursors

    7.2 Spectroelectrochemistry of Catenanes and Their Precursors

    VIII. COMPARISONS WITH MOP AC CALCULATION 164

    8.1 Introduction

    8.2 Performance of the Computations

    8.3 Results of the Computations

    IX. CONCLUSIONS AND SUGGESTIONS FOR FURTHER WORK 170

    9.1 Some Concluding Points

    9.2 Suggestions

    BIBLIOGRAPHY 174

    vu

  • LIST OF TABLES

    PAGE

    Table 3.1 Cyclic Voltammetry Data for Bipyridines 32

    Table 3.2 Cyclic Voltammetry Data for Platinum(II) Complexes 37

    Table 3.3. Spectroscopic Data and Proposed Assignments for Platinum Complexes and

    Related Species in DMF 60

    Table 4.1 Electrochemical data for Ru(Bipy)2(biiq)2+ and related species 77

    Table 4.2 Electrochemical Data for Ruthenium(II)

    Methylsilylmethyl Bipyridine Complexes 95

    Table 4.3. Spectroscopic Data and Proposed Assignments for [Ru(l,r-biiq)(bipy)2]2+

    and Related Species in DMF 96

    Table 4.4. Spectroscopic Data and Proposed Assignments for Ruthenium(Il)

    Methylsilylmethyl Bipyrimidine Complexes 99

    Table 5.1 Cyclic Voltammetry Data for Trimethylsilylmethyl

    Platinum(II) and Palladium(II) Complexes 109

    Table 5.2 Spectroscopic Data and Proposed Assignments for Trimethylsilylmethyl

    Platinum and Palladium Complexes 117

    Table 6.1 Cyclic Voltammetry Results for Phenanthrolines 128

    Table 6.2 Spectroscopic Data and Proposed Assignments

    for Phenanthrolines 136

    Vlll

  • Table 7.1 Electrochemical Data for Catenanes and Their Precursors . . . . . . . . . . 146

    Table 7.2 Spectroscopic Data and Proposed Assignments

    for Catenanes and Their Precursors 154

    IX

  • LIST OF ILLUSTRATIONS

    PAGE

    Fig. 1.1 Energy level diagram for an octahedral complexes containing ligands with K-

    orbitals. From Bryant, G. M.; Fergusson, J. E.; Powell, H. K. J. Aust. J. Chem. 1971,

    24,257 8

    Fig.2.1 The OTTLE cell used in spectroelectrochemical studies (After Song, J-I. PhD.

    Thesis, University of Glasgow, 1989) 14

    Fig.2.2 Some aza-biphenyl compounds which are discussed in Chapter III 17

    Fig.2.3a Platinum(II) complexes which are discussed in Chapter III 18

    Fig.2.3b Platinum(II) complexes which are discussed in Chapter III 19

    Fig.2.4a [Ru(n)(bipy)2(l,l'-biiq)]2+ and related species which aure discussed in Chapter

    IV 20

    Fig.2.4b Ruthenium(II) complexes which are discussed in Chapter IV 21

    Fig.2.4c Ruthenium(II) complexes which are discussed in Chapter IV 22

    Fig.2.5 Platinum(II) and palladium(II) complexes which are discussed in Chapter V. .23

    Fig.2.6 Phenanthrolines discussed in Chapter VI 24

    Fig.2.7a Catenanes and their precursors which are discussed in Chapter VII 25

    Fig.2.7b Catenanes and their precursors which are discussed in Chapter VII 26

    Fig.3.1 Cyclic voltammogram of 2,4'-bipyridine in DMF (supporting electrolyte 0.1 M

    TBAPF6, scan rate 500 mV/s, at room temperature) 29

  • Fig.3.2 Cyclic voltammogram of 2,4'-bipyridine in DMF (supporting electrolyte 0.1 M

    TBAPF6, scan rate 4.0 V/s, at room temperature) 30

    Fig.3.3 Cyclic voltammogram of N'-methyl-2,4'-bipyridine in DMF (supporting

    electrolyte 0.1 M TBAPF6, scan rate 200 mV/s, at room temperature) 33

    Fig.3.4 Cyclic voltammogram of 4,4'-diphenyl-2,2'-bipyridine in DMF (supporting

    electrolyte 0.1 M TBAPF6, scan rate 500 mV/s, at room temperature) 34

    Fig.3.5 The plot of the cathodic current of 4,4'-diphenyl-2,2'-btpyridine against the

    square root of scan rate 35

    Fig.3.6 Cyclic voltammogram of [P^bipy^lJ in DMF (supporting electrolyte 0.1 M

    TBAPF6, scan rate 0.2 V/s at room temperature) 38

    Fig.3.7 Cyclic voltammogram of [Pt^bipy^lJ in DMF (supporting electrolyte 0.1 M

    TBAPF6, scan rate 2 V/s at room temperature) 39

    Fig.3.8 Cyclic voltammogram of [Pt(bipy)(4-NCpy)Cl)]+ in DMF with supporting

    electrolyte, 0.1 M TBAPF6, scan rate 2 V/s, at room temperature 40

    Fig.3.9 Cyclic voltammogram of [Pt(ph2-bipy)Cy in DMF with supporting electrolyte,

    0.1 M TBAPF6; scan rate 0.2 V/s; at room temperature 42

    Fig.3.10 Cyclic voltammogram of [Pt(Me2-bipy)ClJ in DMF with supporting

    electrolyte, 0.1 M TBAPF6; scan rate 0.5 V/s; at room temperature 43

    Fig.3.11 Cyclic voltammogram of Pt(2,4'-bipyOct-H)Cl2 in DMF with supporting

    electrolyte, 0.1 M TBAPF6; scan rate 200 mV/s; at room temperature 44

    Fig.3.12 Cyclic voltammogram of Pt(2,4'-bipyOct)Cl3 in DMF with supporting

    electrolyte, 0.1 M TBAPF6; scan rate at 200 mV/s; at room temperature 45

    XI

  • Fig.3.13 Cyclic voltammogram of [Pt(2,4'-bipyMe-H)(bipy)]2+ in DMF with supporting

    electrolyte 0.1 M TBAPF6 at room temperature, scan rate 2 V/s 48

    Fig.3.14 Cyclic voltammogram of [Pt(2,4'-bipyMe-H)(py)J2+ in DMF with supporting

    electrolyte 0.1 M TBAPF6 at room temperature, scan rate 0.2 V/s 49

    Fig.3.15 Cyclic voltammogram of [Pt(terpy)Cl]+ in DMF with supporting electrolyte

    0.1 M TBAPF6 at room temperature, scan rate 0.5 V/s 50

    Fig.3.16 Cyclic voltammogram of 2,2':6",2"-Terpyridine in DMF with supporting

    electrolyte 0.1 M TBAPF6 at room temperature 51

    Fig.3.17 Cyclic voltammogram of [PtDCMBClJ in DMF with supporting electrolyte

    0.1 M TBAPF6 scan rate 500 mV/s; at room temperature 52

    Fig.3.18 7t-Orbital diagram for biphenyl 56

    Fig.3.19 Electronic absorption spectra of 2,4'-bipyridine and its one electron reduction

    product in DMF (c = 6.6 x 10"4 M) with 0.1 M TBAPF6; parent (dashed line) and

    singly reduced form (solid lines) 57

    Fig.3.20 Electronic absorption spectra of N'-methyl-2,4'-bipyridinium and its one

    electron reduction product in DMF (c = 6.4 x 10"4 M) with 0.1 M TBAPF6; (Solid line

    — parent; dashed line — singly reduced) 58

    Fig.3.21 Electronic absorption spectra of [Pt(2,4'-bipyOctyl)Cl3] and its one electron

    reduction product in DMF (c = 7.8 x 10"4 M) with 0.1 M TBAPF6 (Solid line —

    parent; Dashed line — singly reduced) 59

    Fig.3.22 Electronic absorption spectra of [Pt(2,4'-bipyOctyl-H)ClJ and its one electron

    reduction product in DMF (c = 5.7 x 10"4 M) with 0.1 M TBAPF6 (Solid line —

    xu

  • parent; Dashed line — singly reduced) 64

    Fig.3.23 Electronic absorption spectra of [PtCbipy^lJ and its one electron reduction

    product in DMF (c = 2.3 x 10 3 M) with 0.1 M TBAPF6 67

    Fig.3.24 Electronic absorption spectra of [Pt(bipy)(4-CNpy)Cl]+ and its one electron

    reduction product in DMF (c = 1.2 x 10"3 M) with 0.1 M TBAPF6 70

    Fig.3.25 Electronic absorption spectra of [PtDCMBClJ and its one electron reduction

    product in DMF (c = 2.3 x 10"4 M) with 0.1 M TBAPF6 71

    Fig.4.1 Cyclic voltammogram of l,l'-biisoquinoline in DMF with supporting

    electrolyte 0.1 M TBAPF6 at room temperature; scan rate 0.5 V/s 78

    Fig.4.2 Cyclic voltammogram of l,l'-biisoquinoline in DMF with supporting

    electrolyte 0.1 M TBAPF6 at room temperature, scan rate 4 V/s 79

    Fig.4.3 Cyclic voltammogram of l,l'-biisoquinoline in DMF with supporting

    electrolyte 0.1 M TBAPF6 at room temperature, scan rate 0.5 V/s 80

    Fig.4.4 Cyclic voltammogram of l,l'-biisoquinoline in DMF with supporting

    electrolyte 0.1 M TBAPF6 at room temperature, scan rate 0.2 V/s 81

    Fig.4.5 Cyclic voltammogram of l,l'-biisoquinoline in DMF with supporting

    electrolyte 0.1 M TBAPF6 at room temperature, scan rate 0.5 V/s 82

    Fig.4.6 Cyclic voltammogram of 2,2'-bisquinoline in DMF with supporting electrolyte

    0.1 M TBAPF6 and ferrocene as internal standard at room temperature; scan rate 0.5

    V/s; from 1.0 to -3.0 V 83

    Fig.4.7 Cyclic voltammogram of 2,2'-bisquinoline in DMF with supporting electrolyte

    0.1 M TBAPF6 at room temperature; scan rate 0.5 V/s; from 1.0 to -2.5 V 84

    Xlll

  • Fig.4.8 Cyclic voltammogram of 2,2'-bisquinoline in DMF with supporting electrolyte

    0.1 M TBAPF6 at room temperature; scan rate 4 V/s 85

    Fig.4.9 Cyclic voltammogram of [Ru(biiq)(bipy)2]2+ in DMF with supporting

    electrolyte 0.1 M TBAPF6 at room temperature; scan rate 0.5 V/s 87

    Fig.4.10 Cyclic voltammogram of [Ru(biiq)(bipy)J2+ in CH3CN with supporting

    electrolyte 0.1 M TBAPF6 at room temperature; scan rate 0.5 V/s 88

    Fig.4.11 Cyclic voltammogram of YoungOl in DMF with supporting electrolyte 0.1 M

    TBAPF6 at room temperature; scan rate 0.5 V/s 90

    Fig.4.12 Cyclic voltammogram of Young02 in DMF with supporting electrolyte 0.1 M

    TBAPF6 at room temperature; scan rate 0.5 V/s; with FeCp2 as internal standard . 91

    Fig.4.13 Cyclic voltammogram of Young03 in DMF with supporting electrolyte 0.1 M

    TBAPF6 at room temperature; scan rate 0.5 V/s; with FeCp2 as internal standard . 92

    Fig.4.14 Cyclic voltammogram of Young04 in DMF with supporting electrolyte 0.1 M

    TBAPF6 at room temperature; scan rate 0.5 V/s; with FeCp2 as internal standard . 93

    Fig.4.15 Electronic absorption spectra of l,l'-biisoquinoline and its one electron

    reduction product in DMF (c = 2.2 x 10"4 M) with 0.1 M TBAPF6; parent (solid line)

    and singly reduced form (dashed lines) 96

    Fig.4.16 Electronic absorption spectra of [Ru(biiq)(bipy)J2+ and its one electron

    reduction product in DMF (c = 5.2 x 10"4 M) with 0.1 M TBAPF6; parent (solid line)

    and singly reduced form (dashed lines) 98

    Fig.4.17 Electronic absorption spectra of youngOl and its one electron reduction

    product in DMF (c = 1.3 x 10"3 M) with 0.1 M TBAPF6; parent (solid line) and singly

    xiv

  • reduced form (dashed lines) 101

    Fig.4.18 Electronic absorption spectra of young02 and its one electron reduction

    product in DMF (c = 1.2 x 10"3 M) with 0.1 M TBAPF6; parent (solid line) and singly

    reduced form (dashed lines) 102

    Fig.4.19 Electronic absorption spectra of young03 and its one electron reduction

    product in DMF (c = 1.0 x 10"3 M) with 0.1 M TBAPF6; parent (solid line) and singly

    reduced form (dashed lines) 103

    Fig.4.20 Electronic absorption spectra of young04 and its one electron reduction

    product in DMF (c = 1.2 x 10'3 M) with 0.1 M TBAPF6; parent (solid line) and singly

    reduced form (dashed lines) 104

    Fig.5.1 Cyclic voltammogram of [(bipym)Pt(CH2SiMe3)J in DMF with supporting

    electrolyte 0.1 M TBAPF6; scan rate 0.5 V/s; at room temperature 110

    Fig.5.2 Cyclic voltammogram of [(bipym)Pt(CH2SiMe3)2] in DMF with supporting

    electrolyte 0.1 M TBAPF6; scan rate 2 V/s; at room temperature I l l

    Fig.5.3 Cyclic voltammogram of [(bipym)Pd(CH2SiMe3)J in DMF with supporting

    electrolyte 0.1 M TBAPF6; scan rate 1 V/s; at room temperature 112

    Fig.5.4 Cyclic voltammogram of [(Me3SiCH2)2Pd(bipym)Pt(CH2SiMe3)2] in DMF with

    supporting electrolyte 0.1 M TBAPF6; scan rate 0.5 V/s; at room temperature . . . 115

    Fig.5.5 Cyclic voltammogram of [(Me3SiCH2)2Pd(bipym)Pd(CH2SiMe3)2] in DMF with

    supporting electrolyte 0.1 M TBAPF6; scan rate 0.2 V/s; at room temperature . . . 119

    Fig.5.6 Electronic absorption spectra of [(bipym)Pt(CH2SiMe3)J and its one electron

    reduction product in DMF (c = 4.5 x 10"4 M) with 0.1 M TBAPF6; parent (solid line)

    xv

  • and singly reduced form (dashed lines) 120

    Fig.5.7 Electronic absorption spectra of [(bipym)Pd(CH2SiMe3)J and its one electron

    reduction product in DMF (c = 1.6 x 10"3 M) with 0.1 M TBAPF6; parent (solid line)

    and singly reduced form (dashed lines) 121

    Fig.5.8 Electronic absorption spectra of [(Me3SiCH2)2Pd(bipym)Pt(CH2SiMe3)2] and its

    one and two electron reduction product in DMF (c = 7.5 x 10"4 M) with 0.1 M

    TBAPF6; parent (solid line) and singly reduced form (dashed lines 122

    Fig. 6.1 Cyclic voltammogram of 1,7-phenanthroline in DMF with supporting

    electrolyte 0.1 M TBAPF6 at room temperature; scan rate 0.5 V/s 129

    Fig. 6.2 Cyclic voltammogram of 4,7-phenanthroline in DMF with supporting

    electrolyte 0.1 M TBAPF6 at room temperature; scan rate 0.5 V/s 130

    Fig. 6.3 Cyclic voltammogram of 1,10-phenanthroline in DMF with supporting

    electrolyte 0.1 M TBAPF6 at room temperature; scan rate 0.5 V/s 131

    Fig. 6.4 Cyclic voltammogram of 1 -methyl-1,10-phenanthrolinium in DMF with

    supporting electrolyte 0.1 M TBAPF6 at room temperature; scan rate 0.2 V/s . . . 132

    Fig. 6.5 Cyclic voltammogram of l-methyl-l,10-phenanthrolinium in DMF with

    supporting electrolyte 0.1 M TBAPF6 at room temperature; scan rate 0.2 V/s . . . 133

    Fig. 6.6a Cyclic voltammogram of 7-methyl-4,7-phenanthrolinium in DMF with

    supporting electrolyte 0.1 M TBAPF6 at room temperature; scan rate 0.2 V/s . . . 134

    Fig. 6.6b Cyclic voltammogram of 7-methyl-l,7-phenanthrolinium in DMF with

    supporting electrolyte 0.1 M TBAPF6 at room temperature; scan rate 0.2 V/s . . . 135

    Fig. 6.7 Electronic absorption spectra of 1,7-phenanthroline and its one electron

    xvi

  • reduction product in DMF (c = 4.5 x 10"3 M) with 0.1 M TBAPF6; parent (solid line)

    and singly reduced form (dashed lines) 138

    Fig. 6.8 Electronic absorption spectra of 4,7-phenanthroline and its one electron

    reduction product in DMF (c = 2.7 x 10"3 M) with 0.1 M TBAPF6; parent (solid line)

    and singly reduced form (dashed lines) 139

    Fig. 7.1 Cyclic voltammogram of [BBEPYBIXYCY]4* (JFS08) in Acetonitrile with

    supporting electrolyte 0.1 M TBAPF6 at room temperature; scan rate 1 V/s . . . . 147

    Fig. 7.2 Cyclic voltammogram of [BBIPYXY]4+ (JFS06) in Acetonitrile with

    supporting electrolyte 0.1 M TBAPF6 at room temperature; scan rate 1 V/s . . . . 148

    Fig. 7.3 Cyclic voltammogram of [BBIPYBIXYCY]4* (JFS07) in Acetonitrile with

    supporting electrolyte 0.1 M TBAPF6 at room temperature; scan rate 2 V/s . . . . 149

    Fig. 7.4 Cyclic voltammogram of JFS03 in Acetonitrile with supporting electrolyte

    0.1 M TBAPF6 at room temperature; scan rate 2 V/s 150

    Fig. 7.5 Cyclic voltammogram of JFS04 in Acetonitrile with supporting electrolyte

    0.1 M TBAPF6 at room temperature; scan rate 2 V/s 151

    Fig. 7.6 Cyclic voltammogram of JFS05 in acetonitrile with supporting electrolyte 0.1

    M TBAPF6 at room temperature; scan rate 2 V/s 152

    Fig. 7.7 Electronic absorption spectra of [BBIPYBIXYCY]4* (JFS08) and its two-

    electron reduction product in Acetonitrile with 0.1 M TBAPF6 (c = 6.3 x 10"4 M) 157

    Fig.7.8 Electronic absorption spectra of two-electron reduced [BBIPYBIXYCY]4*

    (JFS08) in Acetonitrile with 0.1 M TBAPF6 (c = 6.8 x 10"4 M) 158

    Fig.7.9 Electronic absorption spectra of [2]catenane (JFS05) and its two-electron

    xvii

  • reduction product in acetonitrile with 0.1 M TBAPF6 (c = 3.5 x 10"4 M) 159

    Fig.7.10 Electronic absorption spectra of [2]catenane (JFS04) and its two-electron

    reduction product in Acetonitrile with 0.1 M TBAPF6 (c = 5.0 x 10"4 M) 160

    Fig.7.11 Correlation between the band at 262 nm and 401 nm of JFS05 in the process

    of reduction 161

    Fig.7.12 Correlation between the band at 262 nm and 615 nm of JFS05 in the process

    of reduction 162

    Fig. 8.1 Individual atomic orbital contributions to the LUMO of co-planar biphenyl.

    (From Mopac Version 6.00 calculation) 167

    Fig.8.2 Individual atomic orbital contributions to the LUMO of phenanthrene. (From

    Mopac Version 6.00 calculation) 168

    xvm

  • CHAPTER I

    INTRODUCTION

    This dissertation is devoted to the characterization of a remarkable and

    interesting family of electrochemically and spectroscopically active supramolecules

    with an azabiphenyl as active moiety. Their electronic structure has been closely

    defined by comparative examination of their redox potentials, and their electronic

    absorption spectra in their original forms and also, where possible, in their reduced or

    oxidized forms. After initial electrochemical characterization by cyclic voltammetry,

    the reduced and oxidized species were generated electrochemically inside an optical

    transparent thin layer electrochemical (OTTLE) cell by controlled potential electrolysis

    and their electronic spectra were taken in situ.

    1.1 Frontier Molecular Orbitals

    In modem chemistry, the electronic structure of a chemical system, an atom, an

    ion or a molecule, is to a good approximation expressed in terms of an array of two-

    electron "orbitals". These electronic orbitals are ordered in energy, and are filled

    progressively by electrons from lower energy to higher, to form the ground state

    electronic configuration. Of these electronic orbitals, the most important orbitals in

    determining the chemistry of a particular chemical system are the HOMO and LUMO.

    The HOMO (Highest Occupied Molecular Orbital) is occupied last when electrons fill

    the available orbitals from lower energy to higher energy. The LUMO (Lowest

    1

  • 2

    Unoccupied Molecular Orbital) is the next higher orbital after the HOMO. HOMO

    and LUMO are collectively called frontier orbitals, because in general, chemical

    processes start with the interaction and reorganization of the frontier orbitals, HOMOs

    and LUMOs, of the participating chemical systems. The chemical properties of a

    chemical system are largely defined by its frontier orbitals.

    By reduction or oxidation, electrons can be added into or removed from a

    molecule. The electron added will go into the LUMO first, and the electron removed

    first will come from the HOMO. The reduction and oxidation potentials which tell

    how easily the chemical system under study can be reduced or oxidized, can be easily

    measured electrochemically. The redox potentials provide measurements of the energy

    levels of the HOMO and LUMO. Electronic spectroscopy is another probe to look

    into the electronic structure of a chemical system. In the electronic transition process,

    an electron in a HOMO or near HOMO orbital gets energy from a photon, and

    promotes itself into an orbital with higher energy, leaving a hole in the original orbital,

    thus forming the excited state. If we neglect changes in electron-electron repulsion

    energy, the position of the bands in the electronic spectra is a measure of the energy

    difference between the two participating orbitals. This is of course a drastic over-

    simplification, but the values obtained may still be useful for comparisons within a

    group of similar molecular species. Because the mass of an electron is much smaller

    than that of a nucleus, and the time scale for an electronic transition is much shorter

    than that for the vibrational relaxation needed by geometry adjustment, the shape of

    the molecule is frozen at its ground state in the electronic transition process, (Franck-

  • 3

    Condon Principle), so that usually the electronic spectra are broad and sometimes

    show vibrational coupling structure. In electrochemical reduction and oxidation,

    however, the species has time to adopt its equilibrium geometry in both oxidation

    states.

    The molecular orbitals are intrinsically delocalized; all orbitals of the atoms

    making up the molecule are more or less fused together by overlapping to form the

    molecular orbital. But in many cases, the localized molecular orbital model is still

    suitable and more descriptive in practice. In the localized molecular orbital model, the

    orbitals can be described as based on each individual fragment For example, in

    coordination compounds, they can be classified as metal-based d orbitals and ligand-

    based a and K orbitals. In this way, the overlapping of the orbitals from different

    parts of the molecule is ignored (zeroth-order approximation). Under this localized

    molecular orbital model, the electronic absorption bands of complexes can be

    classified accordingly into metal-based, ligand-based, and charge transfer bands.

    Transitions between molecular orbitals mainly localized on the metal center are termed

    metal-based transitions, they are generally d -» d transitions for transition metals even

    though some time s -» p and p -¥ d transitions are observed. Transitions between

    molecular orbitals mainly localized on the ligands are termed ligand-based transitions;

    they are mainly % -» 7t* transitions in aromatic ligands such as those under study. In

    other cases a -» a*, n a* and n -> ji* transitions also exist. Transitions between

    molecular orbitals localized at different parts of a molecule which cause the

    redistribution of electronic charge are called charge transfer transitions. More

  • 4

    specifically, in coordination complexes, the charge transfer transitions can be divided

    into metal to ligand charge transfer, MLCT, ligand to metal charge transfer, LMCT,

    and in some cases ligand to ligand charge transfer. MLCT is usually observed when

    there is an oxidizable metal center. Sometimes it is also called as metal oxidation

    charge transfer because it has the metal center oxidized to a higher oxidation state in

    the excited state compared with the ground state. In the same way, LMCT can also be

    called metal reduction charge transfer, and is often observed when the metal center is

    reducible. Under the localized electronic Orbital model, the reduction of a

    coordination compound can be classified as a ligand-based reduction if the added

    electron resides in a ligand-based orbital, or a metal-based reduction if the added

    electron resides in a metal-based orbital; similarly for oxidation.

    The positions of electronic transition bands reflect the energy difference

    between the molecular orbitals involved. The intensity of a band is largely governed

    by the selection rules. For the electronic absorption, which is mainly from an

    electronic dipole transition, the selection rule requires that the system has a transition

    moment between the ground and excited states under the influence of an electric

    dipole so that it can interact with the oscillating electric vector and get energy

    transferred. For an electronic absorption band to have significant intensity, it must

    arise from an allowed transition. Usually some n -» k and many charge transfer

    transitions are orbitally allowed. For an octahedral complex, the metal-based d -» d

    transitions are orbitally forbidden, so they usually are very weak.

    The observation of the changes in redox potential and in the energy of

  • 5

    electronic absorption bands for a given component in different molecular environments

    can be used as a way to investigate the interaction between this component and its

    environment. The electron transfer absorptions are absent from the individual

    components, and directly result from transitions between molecular orbitals localized

    at different components of the supramolecular system.

    1.2 Supramolecules

    The localized molecular orbital theory is the fundamental theory for

    supramolecular chemistry. Supramolecular chemistry studies the chemistry of systems

    (supramolecules) made up of molecular components in the same way as molecules are

    made up of atoms.1 A few examples of supramolecules are catenanes, rotaxanes, and

    donor-acceptor complexes. Most coordination complexes can also fit into this

    category, regarding metal and ligands as the building blocks.

    The term supramolecule is used here to refer to a concept rather than a certain

    category of molecules. The name of supramolecule is given to certain substances, not

    because they are large in size, but because their properties under examination can be

    understood based on the localized electronic theory. The features of a supramolecule

    can be pictured according to its individual molecular components, whose behavior as a

    part of a supramolecule is the direct development from those of the isolated parts or of

    suitable model compounds. It is no surprise that those properties are more or less

    modified when the molecules are organized into a supramolecular system, but in many

    cases they can be understood based on the intrinsic properties of the components,

    considering those modifications which result from interactions within supramolecular

  • 6

    systems as perturbations. The interactions among atoms within classical molecules are

    mainly ionic and covalent interactions. In addition to these, in supramolecular

    chemistry, there are also donor-acceptor, dispersion dipolar, hydrophobic, and

    sometimes hydrogen bonded interactions which are not of primary importance in the

    study of the intermolecular interactions for classical molecules but are important for

    supramolecules.

    1.3 Azabiphenyl Systems

    The chemical systems under study are those containing polypyridine or

    phenanthroline moieties. 2,2'-Bipyridine and 1,10-phenanthroline, as two members of

    this family, have been well known for more than a hundred years as effective

    bidentate ligands. The red-colored ferrous salt of 2,2'-bipyridine was first reported by

    Blau in 1888.2 In 1898, Blau reported the synthesis of 1,10-phenanthroline and

    demonstrated its similarity to 2,2'-bipyridine. He also discovered the reversible nature

    of oxidation of the iron(II) complexes.3 In 1912 Werner demonstrated an octahedral

    configuration for the tris-2,2'-bipyridine iron(II) cation by successful resolution of its

    optical forms.4 In 1928, Manchot and Lehmann used 2,2'-bipyridine to study the

    reaction of ferrous iron with hydrogen peroxide.5 Feigl and Hamburg described a

    similar qualitative application three years later® Apparently, Bode was the first person

    to use the reagent for quantitative purposes, determining iron in beer.7 Widespread

    interest in the analytical applications of these bipyridine and phenanthroline complexes

    developed in 1931 when Hammett, Walden and Chapman described the use of the

    iron(II) complexes as reversible, high potential oxidation-reduction indicators.8

  • 7

    Studies of the crystal structure of bipyridine reveal that the two pyridine rings

    are nearly coplanar with N-atoms in the trans configuration.9 In solution, dipole

    moment measurements indicate that the molecule is approximately planar and also in

    the trans arrangement.10"12 The cisoid form, undoubtedly is adopted for chelate ring

    formation with metal ions; and with properties comparable to phenanthroline, the five-

    membered chelate ring is most probably very close to coplanar with the rest of

    bipyridine molecule.

    The electronegativity of nitrogen is higher than that of carbon, so the n*

    orbitals of bipyridines and phenanthrolines are lower in energy than that of biphenyl

    and phenanthrene, their all-carbon analog, and they can function as good n acceptor

    ligands so as to stabilize low oxidation states of the metal center. These bipyridine

    and phenanthroline complexes are rich in spectroscopy, with bands in the ultraviolet,

    visible and sometimes the infrared region, because of the lower it* orbital of bipyridine

    and phenanthroline. The electronic absorption bands of the complexes directly relate

    to their electronic structure, their frontier orbitals. All of these complexes are also

    electrochemically reducible under suitable experimental conditions.

    1.4 Symmetry Considerations

    Six coordinated tris-complexes of bidentate ligands such as M(bipy)3 or

    M(phen)3 have D3 symmetry. When one or two of these bidentate ligand is substituted

    by other monodentate ligands, the symmetry of the molecule is changed to G, for cis-

    [M(bipy)2XJ, Q for cis-[M(bipy)2XY], D2h for trans-[M(bipy)2XJ, and for trans-

    [M(bipy)2XY]. Mono-bipyridine complexes belong to C,v for [M(bipy)X4], or C[

  • ' l

    \

    (n + l)p / ~2 l̂ \

    —f '1 rm -x

  • 9

    for [M(bipy)X2Y2J. In an Oh environment, if the metal ion coordinates with six

    identical monodentate ligands, the metal-based d orbitals will split into two groups,

    dxy, dyz and dzx with lower energy belonging to t,g and dz2 and dl2.y2 with higher energy

    belonging to eg. There are 12 ligand-Jt orbitals for the octahedral complexes. They

    split into tlg, tlu, tjg and 4 sets of group orbitals in an octahedral field. Among the

    4 sets of group orbitals, tlg, and t2u are nonbonding in character, and tlu interacts with

    (n+l)p. Only will interact with metal center's d(tjg) orbitals to form % bonds. The

    interaction between metal t2g and ligand n* orbitals is an important factor governing

    the chemistry of metal complexes. The simplified energy-level diagram is pictured as

    Fig. 1.1. Tris-bidentate metal complexes have D3 symmetry. There are only three 7t

    orbitals corresponding to the ligand HOMOs, and three for the LUMOs. Their

    chemistry can often be discussed under Oh group symmetry. While doing this, one

    should be aware that the D3 group is only a sub-group of Oh and it is not strictly

    correct to consider the metal d-orbitals as and eg as the degeneracy will be reduced

    by the lower symmetries. Under D3 symmetry the t,g orbitals will split into a,+e,

    while the tlg orbitals split into a ^ e . The energy gaps between those a and e orbitals

    are not large but some believe that they are sources of structure on the intraligand and

    charge-transfer bands of the tris-bipyridine complexes.14,15 Comparing with tris-

    complexes, the lowering of symmetry on going to bis- and mono-bipyridine

    complexes, is even less significant, and their spectra (in solution and room

    temperature) do not differ in band structure to any great extent from those of the tris-

    complexes. The same arguments are also valid for the four-coordinated complexes of

  • 10

    bipyridines. If they adopt planar geometry, they can be treated under D4h symmetry.

  • REFERENCES

    (1) Stoddart, J. F. Chem. Brit. 1988, 24, 1203.

    (2) Blau, F. Ber. 1888, 21, 1077.

    (3) Blau, F. Monatsh, 1898, 19, 647.

    (4) Werner, A. Ber. 1912, 45, 433.

    (5) Manchot, W.; Lehmann, G. Ann. Chem. 1928, 460, 191.

    (6) Feigl, F., Hamburg, H. Z. Anal. Chem. 1931, 86, 7.

    (7) Bode, B. Wochschr. Brau. 1933, 50, 321.

    (8) Walden, G. H. Jr.; Hammett, L.P.; Chapman, R. P. J. Am. Chem. Soc., 1931,

    53, 3908.

    (9) Cagle, F. W. Jr. Acta Cryst. 1948, 1, 158.

    (10) Fielding, P. E.; Lefevre, R. J. W. J. Chem. Soc., 1951, 1811.

    (11) Cumper, C. W. N.; Giaman, R.F.A.; Vogel, A. I. J. Chem. Soc., 1962, 1188.

    (12) Cureton, P. H.; Lefevre, C. G.; Lefevre, R. J. W. J. Chem. Soc., 1963, 1736.

    (13) Bryant, G. M.; Fergusson, J. E.; Powell, H. K. J. Aust. J. Chem. 1971, 24, 257.

    (14) MacCaffery, A. J.; Mason, S. F.; Norman, B. J. J. Chem. Soc. (A), 1969, 1428.

    (15) Ferguson, J.; Hawkins, C. J.; Kane-Maguire, N. A. P.; Lip, H. Inorg. Chem.,

    1969, 8, 771.

    11

  • CHAPTER H

    EXPERIMENTAL

    2.1 Materials

    The solvents used in electrochemistry and spectroelectrochemistry were HPLC

    grade acetonitrile or DMF, purchased from Aldrich. The acetonitrile was distilled

    twice over phosphorus pentoxide, and DMF twice over calcium hydride before use.

    All distillations were carried out under an atmosphere of nitrogen which was dried

    through a molecular sieve column. The freshly distilled solvent was transferred to the

    electrochemistry cell with a syringe. The supporting electrolyte was either tetra-n-

    butylammonium tetrafluroborate or hexafluorophosphate, which were purchased from

    Aldrich and pre-dried in an oven at 120°C for about 2 to 4 hr. before being used.

    Before each set of experiments, the cyclic voltammogram of blank solvent with

    supporting electrolyte was collected to confirm that their quality was satisfactory.

    2.2 Electrochemistry and Spectroelectrochemistry

    The electronic spectra were collected with a Lambda 9 UV-VIS-NIR

    spectrometer made by Perkin Elemer Corporation. The electrochemical experiments

    were performed with an EG & G Princeton Applied Research Potentiostat Model 273

    Potentiostat/Galvanostat, controlled by an IBM compatible computer with

    Electrochemical Analysis Software V 4.01 Beta. Cyclic voltammograms were

    recorded at room temperature, the results being saved as electronic files and displayed

    12

  • 13

    or plotted out when needed.

    Cyclic voltammetry experiments were performed under an atmosphere of argon

    which was dried by passing through a column filled with 4 A molecular sieve and

    then saturated with the appropriate solvent. The same argon gas was also used to

    purge all solutions prior to experimentation. The working concentration of analyte

    was usually around 10"3 M with 0.1 M tetrabutylammonium tetrafluroborate or

    hexafluorophosphate as supporting electrolyte. The routinely used scan rate was 0.5

    V/s, but the results were verified by changing the scan rate over the range 20 mV/s to

    4.0 V/s, which in no case caused significant shift of the peak potentials. The

    electrochemical system used here was a standard three-electrode system. The cell used

    for cyclic voltammetry was a single-compartment-three-electrode cell bought from EG

    & G Company. The working electrode was the cross-section of a platinum wire 0.368

    mm in diameter which was sealed in a glass tube, and the counter-electrode was a

    piece of the same platinum wire about 5 mm in length. The reference electrode was

    Ag/0.01M-AgN03 in a suitable solvent with 0.09 M supporting electrolyte. It was

    connected to the bulk solution through a salt bridge of 0.1 M supporting electrolyte in

    the working solvent with porous Vycor frits. Potentials were reported against the half-

    wave potential for the oxidation of ferrocene, which was added at the end of each run

    as an internal standard; the peak to peak separation for the standard in all cases was

    within the range 60-80 mV. All the glassware and supporting electrolyte used were

    pre-dried in an oven at about 120°C and cooled down to room temperature inside a

    desiccator before being used. The working electrode was polished with sand paper,

  • 14

    and all the platinum electrodes were washed with concentrated nitric acid and then

    rinsed with water and properly dried before being used.

    Spectroelectrochemical experiments were carried out at room temperature.

    When appropriate, the same solution used in cyclic voltammetry was used for

    spectroelectrochemical experiments, sometimes diluted.

    Spectroelectrochemical experiments were carried out in a specially designed

    quartz cell with one millimeter optical path length, as shown in Fig. 2.1. It is also a

    three-electrode electrochemical system with the same reference electrode as used in the

    cyclic voltammetry experiment. The working electrode is a platinum gauze served as

    an Optically Transparent Thin-Layer Electrode, or OTTLE in short, which is mounted

    in the cell across the light beam of the spectrometer. The counter-electrode here is a

    platinum gauze isolated from the bulk solution with a porous Vycor frit to avoid

    contaminating the main cell by its electrochemical products. The spectrum of the

    parent species was collected first, and the potential of the working electrode was then

    set at the desired position and spectra of the reduced species were collected

    continuously, until there were no further changes when the reduction was regarded as

    complete. Extinction coefficients were caCulated for the electrochemically generated

    species assuming quantitative conversion. Generally, the regeneration of parent

    material was followed spectroscopically, and the isosbestic points during reduction

    were also inspected, to check for decomposition or side-reactions. The reference cell

    used while collecting the electronic absorption spectra was a standard 1 mm quartz

    cell, filled with the same solvent with supporting electrolyte used for working solution

  • 15

    Platinum gauze counter electro

    1 mm UV cell

    Ag/Ag+ reference electrode

    orous vycor frits

    Platinum gauze working electrode

    Fig.2.1 The OTTLE cell used in spectroelectrochemical studies (After Song, J-I. PhD.

    Thesis, University of Glasgow, 1989)

  • 16

    and containing a similar piece of platinum gauze for background correction.

    2.3 The Compounds Under Examination

    The structures of the chemical systems under examination are illustrated in the

    following diagrams.

    In Fig. 2.2, 2,2'-bipyridine, 2,4'-bipyridine and 4,4'-diphenyl-2,2'-bipyridine

    were purchased from Aldrich. All platinum(II) complexes shown in Fig. 2.3, were

    kindly supplied by F. Wimmer of Universiti Brunei Darussalam, Bansar Seri Begawan,

    Brunei. In Fig. 2.4, 2,2'-bisquinoline was purchased from Aldrich, and 1,1'-

    bisisoquinoline and [Ru(II)(bipy)2(l,r-biiq)]2+ (l,l'-biiq = l,l'-biisoquinoline) were

    kindly supplied by M. T. Ashby of The University of Oklahoma. The Ru(II)

    complexes in Fig. 2.4 and the platinum(II) and palladium(II) complexes in Fig. 2.5

    were supplied by G. B. Young of Imperial College of Science, Technology and

    Medicine, London, UK. The catenanes and their precursors in Fig 2.7 were supplied

    by J. F. Stoddart of The University, Sheffield, UK.

    2.4 Molecular Orbital Calculations

    Molecular orbital calculations were performed on a UNIX platform with a

    MOP AC package, version 6.00, by Frank J. Seiler Research Laboratory, U.S. Air

    Force Academy. The Hamiltonian used was PM3.

  • 17

    2,2'-bipyridine 2,4'-bipyridine

    N'-methyl-2,4'-bipyridinium N'-methyl-2,2'-bipyridinium

    N'-octyl-2,4'-bipyridinium 4,4'-diphenyl-2,2'-bipyridine

    Fig. 2.2 Some aza-biphenyl compounds which are discussed in Chapter DI

  • 18

    N. el \ / Pt(II)

    CI

    rr Pt(II) /\ N CI

    CN

    [Pt(bipy)Cl2]

    Pt(II)

    [Pt(bipy)(4-NCpy)Cl]-t

    Pt(II)

    [Pt(ph2-bipy)Cl2] [Pt(Me2-bipy)Cl2]

    Pt(II) Pt(II)

    [Pt(2,4' -bipyOc t-H)Cl2]

    N I Oct

    [Pt(2,4'-bipyOct)Cl3]

    Fig. 2.3a Platinum(II) complexes which are discussed in Chapter IE

  • 19

    Pt(II) Pt(II)

    [Pt(2,4'-bipyMe-H)(bipy)]2+ [Pt(2,4'-bipyMe-H)py2P+

    — p t d D i I CI

    [Pt(terp)Cl]+

    Pt(II)

    [Pt(DCMB)Cl2]

    Fig.2.3b Platinum(II) complexes which are discussed in Chapter in

  • 20

    1,1' -biisoquinoline

    2,2'-bisquinoline

    N ii— Ru (11) •

    [Ru(II)(bipy )2( 1,1' -biiq)]2+

    Fig.2.4a [Ru(n)(bipy)2(l,l'-biiq)]2+ and related species which are discussed in Chapter IV

  • 21

    t-Bu P(OCH3)3

    t-Bu

    Ru(II)

    CH2-^.Si

    P(OCH3 ) 3

    [(bipy')Ru[P(OCH3)3]2CH2CH2SiMe3 ] (YoungOl)

    t-Bu

    t-Bu

    P(OCH3)3

    t 4*CH2SIMe3 Ru (II)

    P(OCH3)3

    CH2SIMe3

    [(bipy')Ru[P(OCH3)3]2(CH2SiMe3)23(Young02)

    Fig.2.4b Ruthenium(II) complexes which are discussed in Chapter IV

  • 22

    t-Bu.

    t-Bu

    P(OCH2CH3) 3

    t %%CH2SIMe3 Ru (II]

    P{OCH2CH3)3

    CH2SIMe3

    [(bipy')Ru[P(OCH2CH3)3]2(CH2SiMe3)23(Young03)

    P(OCH2CH3)

    t*CH2SIMe3 Ru(II)

    ^P(OCH2CH3)3

    CH2SIMe3

    [(Me2-bipy)Ru[P(OCH2CH3)3]2(CH2SiMe3)2](Young04)

    Fig.2.4c Ruthenium(II) complexes which are discussed in Chapter IV

  • 23

    n \ yCH;-SiMe:. yCH2-SiMe3

    I / V 1 1 ' I M < n >

    ^ \ / CH2-SiMe, NCH2-SiMe,

    [(bipym)Pt(CH2SiMe3)2] [(bipym)Pd(CH2SiMe3)2l

    r i , ^ N v

    J Me3Si-CH2v / yCH2-SiMe3

    Pd(ll) I .Pt(II) Me3Si-CH2'

    r \ / CH2-SiMe3 NT

    u

    [(Me3SiCH2)2Pd(bipym)Pt(CH2SiMe3)2]

    i f ^ i

    Me3Si'CH2v / S

    yCH2"SiMe3 Pd(II) I p d ( H )

    Me3Si-CH2^ CH2-SiMe3

    U

    [(Me3SiCH2)2Pd(bipym)Pd(CH2SiMe3)2]

    Fig.2.5 Platinum(II) and Palladium(II) complexes which are discussed in Chapter V

  • 24

    1,10-phenanthroline 4,7 -phenanthroline

    1,7 -phenanthroline CH3

    N-methyl-1,10-phenanthrolinium

    7-methyl-1,7-phenanthroline

    Fig.2.6 Phenanthrolines discussed in Chapter VI

  • 25

    JFS-01 JFS-03

    o ^ r c P o ^ „

    JSF-02

    / v t s / s V o o o o o

    r - \ / o o

    6

    + 0

    w U

    0

    A _ / °

    JFS-04

    Fig.2.7a Catenanes and their precursors which are discussed in Chapter VII

  • 26

    JFS-05

    n n n o o o a o o o

    o o o q o o c V - / W W U \ ( w

    JFS-06 JFS-07

    rO~On 8

    JFS-08 ( - H Z Z H - j = c u p ^ ^ ^

    (—c=>—| = ^ ~ V ^ | CH,

    ? - r\J~ f

    Fig.2.7b Catenanes and their precursors which are discussed in Chapter VII

  • CHAPTER m

    ELECTRO- AND SPECTROELECTROCHEMICAL STUDIES OF PLATINUM(II)

    BIPYRIDINE COMPLEXES AND RELATED SPECIES

    Bidentate aza-biphenyl and terphenyl ligands readily form stable complexes

    with platinum(II). Such complexes have been extensively studied in coordination and

    organometallic chemistry.1"20 Some of them also show significant functions in

    biochemistry-related processes. Some cis-platinumammine compounds act as

    antitumor drugs by binding covalently to DNA under specific conditions.21 The

    preparation of platinum(II) 2,2'-bipyridine complexes was first reported by Morgan

    and Burstall about sixty years ago.1 2,2'-Bipyridine is an important bidentate ligand

    which forms complexes with most of the transition metals. It also can be converted to

    a monodentate ligand by quaternizing one-of the nitrogens, resulting in a cation which

    is isoelectronic with 2-phenylpyridine.22"26 The resulting cation can undergo ortho-

    metallation at the C(3) position of the ring being quatemized to become the bidentate

    zwitterionic ligand (Mebipy - H) which is isoelectronic with bipyridine.27'28 All of

    these platinum(II) complexes have very rich electrochemistry and informative spectra.

    Their electrochemical and spectral properties have been investigated in detail.

    2,4'-Bipyridine differs from 2,2'-bipyridine as its two nitrogens are located at

    unsymmetrical positions, and it can not serve as a chelating bidentate ligand with two

    nitrogen atoms as donors bonding with one metal center. But when it coordinates

    27

  • 28

    through the nitrogen on the 2-pyridine ring, it can undergo orthometallation at the C(3)

    position of the 4-pyridine ring, forming complexes with an anionic ligand which is

    isoelectronic with the bipy-H anion. By quaternizing on the 4-pyridine ring, we can

    make a cationic species which can coordinate with metal ions as the same way as

    above, but as a zwitterionic bidentate ligand isoelectronic with the bipyridines.

    3.1 Electrochemistry of Bipyridines and Platinum(II) Complexes

    3.1.1 Electrochemistry of Free Bipyridines

    The electrochemistries of 2,2'-bipyridine and 4,4'-bipyridine have been

    investigated before,29 but there are no reports concerning 2,4'-bipyridine.

    3.1.1.1 Electrochemistry of 2,4'-Bipyridine

    The cyclic voltammogram of 2,4'-bipyridine as shown in Fig.3.1 and Fig.3.2

    has two accessible reductions in DMF. The first one at -2.50 V (vs. FeCp2 /FeCp2+) is

    a chemically reversible one electron reduction. As with 2,2'-bipyridine,29 the second

    reduction of 2,4'-bipyridine is chemically irreversible at the scan rate up to 4 V/s.

    Brown and Butterfield,32 explained this by suggesting that the electrochemically

    formed dianion extracts a proton from the tetrabutylammonium cation leading to its

    own degeneration and the formation of butene and tributylamine. In the biphenyl

    system, the 7t(7) orbital has larger electron densities at para than at ortho positions, so

    the nitrogen, whose electronegativity is higher than carbon, will stabilize the 7t(7)

    orbital better when it is at a para position than when at an ortho position. That is why

    the potential of the first reduction of 2,4'-bipyridine lies between the first reduction

    potential of 2,2'-bipyridine (-2.56 V) and 4,4'- bipyridine29 (-2.40 V) under the same

  • 29

    1 I I 1 o o o o o o o o o o © o

    cu © OJ 1

    S

    tu

    (Vn) i

    Fig.3.1 Cyclic voltammogram of 2,4'-bipyridine in DMF (supporting electrolyte 0.1 M

    TBAPF6, scan rate 500 mV/s, at room temperature)

  • 30

    » SI 5 2

    o 8

    o o

    OJ I

    0 3 c\i 1

    0 8 cu 1

    o

    i i

    m i

    § i o s

    o o o o o o

    O O O I ?

  • 31

    experimental conditions. The reduction potentials of bipyridines are collected in Table

    3.1.

    3.1.1.2 Electrochemistry of N'-Methyl-2,4'-bipyridine

    The cyclic voltammogram of the mono-quaternized 2,4'-bipyridine, N'-methyl-

    2,4'-bipyridinium, in Fig.3.3, also shows two accessible reductions, and in contrast to

    2,4'-bipyridine, both of them are chemically reversible. The doubly reduced species is

    stabilized in this ionic case. It seems that the positive charge introduced by

    quaternization slows down the protonation process which happens on the doubly

    reduced bipyridines. The reduction of quaternized bipyridines is easier than that of the

    parents, a result which can obviously be attributed to their positive charge. The

    potentials of both first and second reductions shift about one volt positively on

    quaternization of 2,4'-bipyridinium, as previously reported for the 4,4'-analog.29

    3.1.1.3 Electrochemistry of 4,4'-Diphenyl-2,2'-Bipyridine

    Fig 3.4. shows the cyclic voltammogram of 4,4'-diphenyl-2,2'-bipyridine. It

    have two accessible reductions in DMF. As in the other bipyridines, the first

    reduction is reversible as told by the current ratio of cathodic and anodic waves being

    close to unity. The plot of its cathodic current against the square root of the scan rate

    is a straight line over the range of scan rate from 0.1 to 4 V/s, as shown in the Fig

    3.5, telling that the electrode process controlled by diffusion.40 The potential of this

    reduction is at -2.42 V, slightly less negative than that of 2,2'-bipyridine. This may

    result either from an inductive effect of the electron-withdrawing phenyl rings, or from

    derealization. The second reduction is chemically irreversible, as in the other

  • 32

    Table 3.1 Cyclic Voltammetry Data for

    Bipyridines"

    Compounds E°/-i E-i/-2

    2,4'-bipyridine -2.50/122" -3.21/irr°

    N'-methyl-2,4'-bipyridine -1.52/92 -2.32/79

    4,4'-diphenyl-2,2'-bipyridine -2.42/93 Not observed

    2,2-bipyridine29 -2.56/74 -3.20/irr

    4,4'-bipyridine29 -2.40/69 -2.88/irr

    aData taken from cyclic voltammetry at 500 mV/s; measurements taken in V vs.

    oxidation potential of ferrocene as internal standard, in dry DMF, at room temperature

    with 0.1 M TBAPF6 as supporting electrolyte.

    b ^(Epa+Epc)(V)/(Epa-Epc)(mV)

    c Chemically irreversible at 4 V/s

  • 33

    o to

    S UJ

    (Vn) i

    Fig.3.3 Cyclic Voltammogram of N'-methyl-2,4'-bipyridine in DMF (supporting

    electrolyte 0.1 M TBAPF6, scan rate 200 mV/s, at room temperature)

  • 34

    o o r*N.

    o o o o o o

    o o o o o o

    o o o o o o

    o o o o o o

    o © o

    to in m cu

    (Vn)

    VI

    I

    o 1

    cu 1

    o 8

    O o

    ai i

    o 8 > C\l I

    0 s 01 t

    o 0 rv oj 1

    o s

  • 35

    cn GO co in

    Fig.3.5 The plot of cathodic current of 4,4'-diphen-2,2'-bipy against the square root of

    scan rate

  • 36

    bipyridines.

    3.1.2 Electrochemistry of Platinum(II) Complexes

    In general, the platinum(II) complexes investigated here show two or three

    single electron reductions but no accessible oxidation under our experimental

    conditions. Most of the reductions are chemically reversible as shown by their cyclic

    voltammograms (potentials summarized in Table 3.2), but the separation of forward

    and return waves is larger in most cases than the ideal value of 59 mV; in other

    words, they are not completely electrochemically reversible. According to the

    localized molecular orbital model, the reductions of these platinum(II) complexes can

    be classified as ligand-based or metal-based.

    3.1.2.1 Ligand-based Reductions of Platinum(II) Complexes

    Bipyridines are well known for their low lying % orbitals. They usually serve

    as K acceptor ligands to form complexes with a wide variety of metals. Most

    complexes of this kind have a bipyridine-based LUMO and their first reduction is

    bipyridine ligand-based.30"32 All the complexes studied here have one or more

    bipyridine ligands. It is to be expected that their first reductions will be ligand-based.

    [Pt(bipy)Cl2]

    [Pt(bipy)CL>] has two accessible reductions in the solvent window of DMF as

    shown in its cyclic voltammogram in Fig.3.6 and Fig.3.7. The first reduction is

    chemically reversible, but the separation of the forward and return wave is 90 mV,

    showing some electrochemical irreversibility. The potential of the first reduction is

    -1.63 V, which is 0.93 V less negative than that of free 2,2'-bipyridine29. It is typical,

  • 37

    Table 3.2 Cyclic Voltammetry Data for

    Platinum(II) Complexes11

    E%' E-iw E-2/-3

    [Pt(bipy)ClJ -1.63/90" -2.39/irf

    [Pt(bipy)(4-Ncpy)Cl]+ -1.47/90 -1.69/100 -2.29/100

    [Pt(ph2-bipy)ClJ -1.59/68 -2.18/79

    [Pt(Me2-bipy)Cl2] -1.64/68 -2.38/irf

    [Pt(2,4'-bipyOct-H)ClJ -1.54/90 -2.23/90

    [Pt(2,4'-bipyOct)Cl3] -1.45/90" -2.04/90

    [Pt(2,4'-bipyMe-H)(bipy)]2+ -1.22/80 -1.57/70 -1.96/80

    [Pt(2,4'-bipyMe-H)pyJ2+ -1.44/60 -2.09/130

    [Pt(DCMB)Cl2] -1.16/76 -1.81/100

    [Pt(Mebipy-H)bipy]2+22 -1.46/64 -1.69/62

    [Pt(Mebipy-H)pyJ2+ 22 -1.58/64 -2.12/75

    "Data taken from cyclic voltammetry at 500 mV/s; measurements taken in V vs.

    oxidation potential of ferrocene as internal standard, in dry DMF, at room temperature

    with 0.1 M TBAPF6 as supporting electrolyte. b Vi(Epa+Epc)(V)/(Epa-Epc)(mV) 0 Chemically irreversible at 4 V/s

  • 38

    o o cu

    o o CXI

    «

    0 1

    o o to * 0 1

    o o o

    I ^ >

    o o o o o o o o o o in in in in tn in cn OJ

    o o in

    o o in •

    0 1

    o o

    o o CD

    O c8 CXI

    o o CO

    in

    UJ

    (vn) i

    Fig.3.6 Cyclic Voltammogram of [Pt(bipy)ClJ in DMF (supporting electrolyte 0.1 M

    TBAPF6, scan rate 0.2 V/s at room temperature)

  • 39

    s O CVJ

    (V«) I

    Fig.3.7 Cyclic Voltammogram of [Pt(bipy)ClJ in DMF (supporting electrolyte 0.1 M

    TBAPF6, scan rate 2 V/s at room temperature)

  • 40

    (vn) i

    Fig.3.8 Cyclic Voltammogram of [Pt(bipy)(4-NCpy)Cl)]+ in DMF with supporting

    electrolyte, 0.1 M TBAPF6, scan rate 2 V/s, at room temperature

  • 41

    when bipyridine is coordinated with a metal ion, that its reduction potential shifts

    positively. This is due to the stabilization of rc(7) orbital of bipyridine by the

    positively charged metal ion. The same thing happens when bipyridines are

    quaternized, as discussed above.29 Klonger, Huffman and Kochi reported the first

    reduction of this complex (by cyclic voltammetry in acetonitrile) in 1982. They

    described the reduction product as a Pt(I) species, but this assignment seems to have

    been made in a formal sense only. The reduction potential of coordinated bipyridine

    is also affected by the ligands coordinated at other sites of the same metal center, due

    to the change of charge distributions. In the case of [Pt(bipy)(py)J2+,22 the first

    reduction, which is also bipyridine-based, happens at -1.38 V, 250 mV less negative

    than the dichloro case, because this is a reduction on a dicationic species. This may

    be attributed to the positive charge and/or -the change from a n donor ligand into a jc

    acceptor ligand at the two other coordination sites, which stabilizes not only the

    LUMO of the metal but also the LUMO of bipyridine; in the case of [Pt(bipy)ClJ, the

    n(7) orbital of bipyridine is destabilized by the negatively charged chloride through L

    —> M —> L' interaction.

    [Pt(bipy)(4-NCpy)Cl]+

    The cyclic voltammogram of [Pt(bipy)(4-NCpy)Cl]+ shows, as in Fig.3.8, three

    reduction waves. The first one, at -1.47 V, is again a bipyridine-based reduction. The

    potential of this reduction is about 160 mV less negative than that of [Pt(bipy)ClJ,

    due to charge and/or the jc acceptor effect of 4-cyanopyridines as against the K donor

    effect of chlorides. The second reduction at -1.69 V can be assigned as 4-

  • 42

    o o o o o o OJ CD

    CVJ

    1 1 I i i

    o o o o o o o o o o o CO cu CM CO

    •1 o o o o 1 *

    (vn) I

    Fig.3.9 Cyclic Voltammograms of [Pt(ph2-bipy)Cl2J in DMF with supporting

    electrolyte, 0.1 M TBAPF6; scan rate 0.2 V/s; at room temperature

  • 43

    o o

    o o o

    o o

    o o 03

    #

    0 1

    o o ^ OJ

    7 u.

    © o CO

    o o 0 OJ 1

    CXI I

    o o 00

    o o cu

    o o 00

    o § O o

    o o o

  • 44

    o m

    C\J cu CU *"• «r1 O I

    (vn) I

    F ig3 . l l Cyclic Voltammograms of Pt(2,4'-bipyOct-H)Cl2 in DMF with supporting

    electrolyte, 0.1 M TBAPF6; scan rate 200 mV/s; at room temperature

  • 45

    o o *

  • 46

    cyanopyridine-based.

    [Pt(ph2-bipy)ClJ and [Pt(Me2-bipy)Cl2]

    As shown in Fig.3.9 and Fig.3.10, the first reduction of [Pt(ph2-bipy)ClJ is at

    -1.59 V and that of [Pt(Me2-bipy)Cl2] is at -1.64 V. They are both bipyridine-based.

    This potential is almost the same for [Pt(bipy)ClJ and [Pt(Me2-bipy)Cy, but shifts

    positively for [Pt(ph2-bipy)Cy due to the derealization or inductive effect of the

    electron-withdrawing phenyl rings.

    Pt(2,4'-bipyOct-H)Cl2 and [Pt(2,4'-bipyOct)Cl3]

    For Pt(2,4'-bipyOct-H)Cl2, the first reduction, as shown in Fig.3.11, is at -1.54

    V, very close to that of N'-methyl-2,4'-bipyridinium. We ascribe this to two opposite

    effects. Coordination with platinum(II) will make the reduction of bipyridine easier,

    but the replacement of a carbon-hydrogen bond by a carbon-platinum bond creates a

    formal negative charge on that carbon atom, making the reduction more difficult.

    Unlike Pt(2,4'-bipyOct-H)Cl2, [Pt(2,4'-bipyOct)Cl3] has only its nitrogen site

    coordinating with platinum(II), without forming a carbon-platinum bond, and is a

    zwitterionic overall neutral complex. Its first ligand-based reduction is at -1.45 V, as

    shown by its cyclic voltammogram in DMF (Fig.3.12), about 90 mV less negative than

    Pt(2,4'-bipyOct-H)Cl2.

    [Pt(2,4' -bipyMe-H) (bipy)]2+

    In its cyclic voltammogram (Fig.3.13), [Pt(2,4'-bipyMe-H)(bipy)]2+ shows three

    reduction waves, of which the first two are based on the two separate ligands. Since

    [Pt(2,4'-bipyOct-H)Cl2] is more readily reduced than [Pt(bipy)Cy, the first reduction

  • 47

    of this complex can be attributed to the 2,4'-bipyMe-H ligand and the second to the

    bipyridine. The reduction potential of 2,2'-bipyridine is at -1.57 V here, which is less

    negative than that of [Pt(bipy)ClJ due to the overall charge on the complex and the it-

    acceptor effect of r-Methyl-2,4'-bipyridine-3-ylium.

    [Pt(2,4'-bipyMe-H)py2]2+

    In [Pt(2,4'-bipyMe-H)py2]2+ (Fig.3.14), the first reduction at -1.44 V, on 2,4'-

    bipyMe-H, is 220 mV more negative than in [Pt(2,4'-bipyMe-H)(bipy)]2+, due to the

    weaker Tt-accepting ability of pyridine compared with bipyridine. The first reduction

    of [Pt(bipy)pyJ, being bipyridine-based, is at -1.38 V, which is less negative than that

    of the 2,4'-bipyMe-H analog by 60 mV. Comparing at the first reductions of [Pt(2,4'-

    bipyMe-H)Cy and [Pt(bipy)ClJ, one finds that the complex of 2,4'bipyMe-H is more

    easily reduced. This contrast can be understood by considering that chlorine is a

    negatively charged % donor, but pyridine is a neutral % acceptor. In the chlorine case,

    the metal center carries less positive charge, but will still interact more strongly with

    2,4'-bipyMe-H than with bipyridine. Pyridine, as a % acceptor, will increase the metal

    center's positive charge, strengthening the interaction between the metal center and

    either bipyridine or 2,4'-bipy-Me-H.

    [Pt(Terpy)Cl]+

    [Pt(Terpy)Cl]+ has three accessible reductions. Its cyclic voltammogram and

    that of free terpyridine are presented in Fig.3.15 and Fig.3.16. The first two

    reductions of the complex are terpyridine-based, and are chemically reversible. Free

    terpyridine shows a chemically reversible reduction at -2.52 V, and the second

  • 48

    o

    s o

    s o

    8 o

    s? o §

    o

    s o

    s o o i n

    O €VJ

    8 ' I D v c o OJ *•1 o o * * CM

    m i

    Fig.3.13 Cyclic Voltammogram of [Pt(2,4'-bipyMe-H)(bipy)]2+ in DMF with supporting

    electrolyte 0.1 M TBAPF6 at room temperature, scan rate 2 V/s

  • 49

    in (xj £

    UJ

    (V«) I

    Fig.3.14 Cyclic Voltammogram of [Pt(2,4'-bipyMe-H)(py)2]2+ in DMF with supporting

    electrolyte 0.1 M TBAPF6 at room temperature, scan rate 0.2 V/s

  • 50

    o o o o o o o o o o c u i n i n i n i n i n i n i n i n i n i n i e u r ^ CM CVJ r ^ c u CU r ^ CU

    c n OJ r u

  • 51

    o 0

    o 1

    o o GO

    »

    0 1

    o o cu

    o § >

    UJ

    o o 0 a] 1

    o o

    cu I

    0 co cvi 1

    CO ai

    CVn) i

    Fig.3J6 Cyclic Voltammogram of 2,2':6",2,,-Terpyridine in DMF with supporting

    electrolyte 0.1 M TBAPF6 at room temperature

  • 52

    o o CXI

    o o CO 0 1

    o o o

    o ©

    tu

    o §

    c3 eg i

    to © £ s

    fv O 8

    o ID % 8 8 o in cu o m rv o ru in i cu

    CO CM OJ

  • 53

    reduction with cathodic wave potential at -2.98 V, for which an anodic wave could

    not be detected even with a scan rate of 4 V/s. As in other cases, the reversibility of

    the second reduction of terpyridine is improved on coordination to the metal ion. The

    first reduction of [Pt(terpy)Cl]+ is at -1.09 V. The first reduction potential of free

    terpyridine, -2.56 V, is close to the value of 2,2'-bipyridine, -2.52 V. But, perhaps

    because there are three nitrogens coordinating with platinum(II) ion, constraining the

    ligand to be planar, the positive shift of reduction potentials of terpyridine due to

    coordination is larger than in the other cases discussed here.

    [Pt(DCMB)Cy

    [Pt(DCMB)ClJ has its first reduction at -1.16 V, which is DCMB ligand-based.

    The cyclic voltammogram of [Pt(DCMB)ClJ is presented in Fig.3.17. It is not

    surprising that the electron-withdrawing carboxymethyl group greatly lowers the

    LUMO of the bipyridine.

    3.1.2.2 Platinum(13)-Based Reductions

    Platinum(II) complexes can undergo metal-based reduction, forming

    platinum(I).29 In the complexes investigated here, most were found to have their

    metal-based reduction located at around 2.0 V, as expected by analogy29 with the

    earlier work. The data are collected in Table 3.2. The metal-centered reduction of

    [Pt(bipy)ClJ shows some chemical irreversibility. In its cyclic voltammogram, at the

    scan rate of 0.5 V/s, there is only an anodic wave centered at -2.39 V, but no

    corresponding cathodic wave. The return wave for the second reduction can be

    resolved at a scan rate of 2 V/s, as shown in Fig.3.7. The potential of the second

  • 54

    reduction for this complex is quite negative, which is mainly due to the % donor effect

    of the dichlorine ligands. Or in other words, this is a reduction being carried out at a

    complex with three negatively charged ligands (one reduced bipyridine and two chlo-

    rides). In contrast, the platinum(II) centered reduction of [Pt(DCMB)ClJ is reversible

    at -1.81 V, about 550 mV less negative than that of [Pt(bipy)ClJ. This is because

    3,3'-dicarboxymethyl-2,2'-bipyridine is presumably a better n acceptor and/or poorer

    a-donor than bipyridine, its LUMO being about 500 mV lower than that of bipyridine,

    due to the electron withdrawing effect of the carboxymethyl group discussed earlier.

    The metal centered reduction of [Pt(2,4'-bipyOct-H)ClJ is 160 mV less negative than

    in [Pt(bipy)ClJ for the same reason; the lower LUMO of 2,4'-bipyOct-H makes it a

    better it acceptor than bipyridine. The second, metal-based reductions of [Pt(2,4'-

    bipyMe-H)pyJ+ and [Pt(bipy)pyJ2+ are almost at the same potential (-2.09 V and

    -2.07 V22), even though their first, ligand-based reductions show that [Pt(2,4'-bipyMe-

    H)py2] is more easily reduced (-1.44 V and -1.38 V22). The metal centered reduction

    of [Pt(bipy)(4-CNpy)Cl]+ appears as the third wave in its cyclic voltammogram at

    -2.29 V, which is much more negative than in other similar complexes. As mentioned

    above, its first reduction is bipyridine-based and the second reduction is 4-cyano-

    pyridine-based. So in this case the platinum(II) ion being reduced is coordinated to

    three negative charged ligands, which causes a negative shift of its reduction potential.

    3.2 Spectroelectrochemistry

    3.2.1 2,4'-Bipyridines

    The electronic absorption spectral data and proposed assignment for bipyridines

  • 55

    are collected in Table 3.3.

    As shown in Fig.3.19, in DMF, the electronic absorption spectrum of neutral

    2,4'-bipyridine shows only one strong band, at 273 nm, which is assigned to the TC(6)

    to TC(7) transition. Bands with higher energy are not observable here due to strong

    absorption by the solvent. The singly reduced 2,4'-bipyridine anion radical generated

    inside the OTTLE cell gives a spectrum having the same features as that of 2,2'-

    bipyridine.29 The strongest band at 388 nm is the transition rc(6) to k ( 7 ) . This band

    shifts from 273 nm in the parent bipyridine to lower energy in the reduced radical

    anion. This kind of shift, which is common29 in the bipyridines, is due to the added

    electron on the %(7) which increases the bond order between the two pyridine rings;

    Jt(7) is an antibonding orbital for the molecule as a whole but bonding between the

    two rings36 (see Fig. 3.18). The bands spreading from 550 nm to 800 nm are assigned

    to the transition from rc(7) to 7t( 10), and bands at 970 nm and 1128 nm to jc(7) -»

    3i(8,9). Both these transition bands have clear vibrational coupling structure with

    separation of 1,400 cm'1. The spectra from quaternized r-methyl-2,4'-bipyridinium in

    the same experimental conditions are similar to those of 2,4'-bipyridine, as in Fig.3.20.

    The 7t(6) to jc(7) transition is at 291 nm in the parent, moved to lower energy due to

    the positive charge, and at 369 nm in the singly reduced species. The k( 7) to 7t(10)

    transition appears at 523 nm, and n{7) to Jt(8,9) in the 1000 nm region.

    3.2.2 Spectroscopy and Spectroelectrochemistry of Square Planar Platinum(II)

    Complexes with 2,4'-Bipyridines as Ligands

    The electronic spectra of these platinum(II) bipyridine complexes can be

  • 56

    -TT 8 u

    00 , , u

    4u

    Fig. 3.18 7C-orbital diagram for biphenyl (After Song, J-I. PhD Thesis, University of

    Glasgow, 1989)

  • 57

    2.40 r

    800 1000 1200

    Wavelength (nm)

    1400 1600

    Fig.3.19 Electronic absorption spectra of 2,4'-bipyridine and its one electron reduction

    product in DMF (c = 6.6 x 10^ M) with 0.1 M TBAPF6; parent (dashed line) and

    singly reduced form (solid lines)

  • 58

    2.20

    8 1.32

    1000 1200 Zl

    1400

    Wavelength (nm)

    Fig.3.20 Electronic absorption spectra of N'-methyl-2,4 -bipyridmium and its o

    , • n u c fc - 6 4 x lO^M) with 0.1 M TBAPF6; parent electron reduction product m DMr (c

    (solid line) and singly reduced form (dashed line)

  • 59

    u o c CS X! h< O 5/3 £> <

    2.80

    2.236

    -0.02 260 400 600

    • • I

    800 1000 1200 Wavelength (nm)

    Fig.3.21 Electronic absorption spectra of [Pt(2,4'-bipyOctyl)Cl3j and its one electron

    reduction product in DMF (c = 7.8 x 10"4 M) with 0.1 M TBAPF6 (Solid line -

    parent; Dashed line ~ singly reduced)

  • 60

    Table 3.3. Spectroscopic Data and Proposed Assignments

    for Platinum Complexes and Related Species in DMFa

    Comp. JC(6) TT(7) Tt(7) JC(10) rt(7) JI(8,9) MLCT Other(see text)

    I" 273(36.6)(20.I)

    I 388(25.8)(36.7) 580(17.2)(11.6) 970(10.3X0.54) 1128(8.9X0.54)

    II 290(34.5X21.2)

    n 368(27.1)(33.4) 480(20.8)sh 524(19.1)(7.35)

    i n 300

    m 347(28.8)(17.9) 375(26.6)(17.5)

    665(15.0X8.7)

    IV 313(31.9X7.9) 311(32.1X6.7)

    278(36.0)(18.9) 388(25.7)(8.9) 363(27.5)sh

    IV 359(27.9)(10.7) 463(21.6X5.1) 913(10.1X1.5) 424(23.6X5.3)

    V 311(32.2)(13.9) 323(31.0X16.6)

    - 333(30.0)sh

    V 273(36.6)

    VI 338(29.6)(12.9) 400(25.0X7.8) 297(33.7)(47.5)

    VI 367(27.2)(34.9) 430(23.3X8.5) 1117(8.9X4.8) 1279(7.8)(4.4)

    524(19.1)(5.0) 562(17.7X5.7)

    VII 322(31.1)(7.7) 311(32.1)(6.7)

    384(26.0)(3.39) 363(27.5)sh 280(35.7X17.7)

    VII 358(27.9)(5.3) 474(21.1)(2.4) 632(15.9)(0.35) 700(14.3X0.34)

    510(19.6X2.8) 418(23.9)(2:.4)

  • 61

    Table 3.3.(Cont.) Spectroscopic Data and Proposed Assignments

    for Platinum Complexes and Related Species iin DMP

    Comp. re(6) -> jt(7) Jt(7) TC(10) rc(7) (8,9) MLCT Other(see text)

    VlIIb 295(33.9X9.3) 268(37.3X11.4)* 372(26.9)sh 432(23.1)(1.2) 451(22.2)(1.2)

    343(29.2)sh

    v i n 319(31.3)(5.4) 473(21.1X2.7) 887(11.3X0.35) 375(26.7)(2.2)' 260(38.5) 538(18.6X2.2) 580(17.2)sh 635(15.7X0.89) 696(14.4)(0.71)

    VIII2 319(31.3X5.4) 393(25.4)(4.3)

    XIII 373(26.8) 436(22.9)

    Xffl 450(22.2) 900(11.1)

    IX

    IX 371(26.9X25.2) 525(19.0X8.2)

    X 312(32.1X20.2) 373(26.8X7.6)

    XI 305(32.8X9.1) 316(31.6)(9.2)

    359(27.8X3.6)

    bipy 281(35.6)

    bipy" 397(25.2)(19.1) 582(17.2)(6.1) 882(11.3X2.6)

    [Pt(bipy)pyJ2+ 22 306(32.7)(18.5) 245(40.8)(10.4)

    [Pt(bipy)py2]+ 22 348(28.7X8.4) 492(20.3)(6.8) 1090(9.2X3.3) 400(25.0)(7.2)

    f Pt(bipy)(Me2N-py)Ji+ a 314(31.8)(17.8) 250(40.0)(9.2)

    [Pt(bipy)(en)]2+ 22 321(31.2X18.4) 248(40.3)(9.0)

  • 62

    Table 3.3, Footnotes and Key

    aData presented as wavelength(A,)/nm(wavenumber(v)/103 cm^Xe/lO3 M"1 cm"1)

    b I, 2,4'-bipyridine, II, N'-methyl-2,4'-bipyridinium,

    III, 4'-diphenyl-2,2'-bipyridine IV, Pt(bipy)Cl2

    V, [Pt(bipy)(4-NCpy)Cir VI, Pt(ph2-bipy)Cl2

    Vn, Pt(Me2-bipy)Cl2 Vffl, Pt(2,4'-bipyOct-H)Cl2

    IX, [Pt(2,4'-bipyOct)Cl3] X, [Pt(2,4'-bipyMe-H)(bipy)]2+

    XI, [Pt(2,4'-bipyMe-H)(py)2]+ Xn, [Pt(terpy)Cl]+

    XIH, Pt(DCMB)Cl2

    bipy, 2,2'-bipyridine

    4-NCpy = 4-cyanopyridine;

    Mej-bipy = 4,4'-dimethyl-2,2'-bipyridine;

    ph2-bipy = 4,4'-diphenyl-2,2'-bipyridine;

    2,4'-bipyOct-H = N'-octyl-2,4'-bipyridin-3'-ylium;

    2,4'-bipyOct = N'-Octyl-2,4'-bipyridinium;

    2,4'-bipyMe-H = N'-Methyl-2,4'-bipyridin-3'-ylium;

    terpy = 2,2':6',2"-terpyridine;

    DCMB = 3,3'-dicarboxymethyl-2,2'-bipyridine

    MejN-py = 4-(dimethylamino)pyridine

    en = 1,2-diaminoethane

  • 63

    assigned with the help of those the free ligands, as summarized in Table 3.2 and 3.3.

    Pt(2,4-bipyOctyl)Cl3:

    The electronic absorption spectra of [Pt(2,4'-bipyOctyl)Cl3] are presented in

    Fig.3.21. The spectrum of the parent is simple. There are bands at high energy that

    overlap with strong absorption background of solvent around 280 nm. These bands may

    be ligand-based 71(6) to rc(7) and/or MLCT, and also some weak MLCT bands in the 400

    nm region. The spectrum of the singly reduced species here is similar to those of

    bipyridine anion radicals. There is a sharp strong band at 371 nm, which is the typical

    rc(6) to Tt(7) transition band of bipyridine anion radicals, and a band at 525 nm which is

    the ti(7) to 7i(10) transition of reduced bipyridinium. All these bands are located at

    almost the same position as in reduced 1 '-methyl-2,4'-bipyridinium. These spectra clearly

    imply that the first reduction of this complex is ligand-based, so that the added electron

    is mainly localized at bipyridine part of complex, as is usual29 in such complexes.

    Another band around 700 nm can be assigned as a ligand-based transition from 7t(7) to

    7t(8,9). This band is, as usual, extremely weak, due to the cancellation of local transition

    dipole moments.

    [Pt(2,4'-bipyOctyl-H)Cl2]

    Compared with Pt(2,4'-bipyOctyl)Cl3, the spectrum of this compound is much

    more complicated and informative as shown in Fig.3.22. The reason is that in the first

    case, the platinum ion is only bonded at one end of the bipyridine ligand, so there is only

    one metal-ligand o bond but no 7t interaction between them. In the second case, the

    situation is totally different.

    In the parent spectrum, there are two well separated bands at 268 nm and 295 nm.

  • 64

    v g 1.176 x> 0

    1 0.970

    0.764 -

    0.362 -

    0.146 -

    600 800 1000

    Wavelength (nm)

    1300

    Fig.3.22 Electronic absorption spectra of Pt(2,4'-bipyOctyl-H)CI2 and its one electron

    reduction product in DMF (c = 5.7 x 10"4 M) with 0.1 M TBAPF6 (Solid line -

    parent; Dashed line — singly reduced).

  • 65

    The intraligand rc(6) —» it(7) transitions of unreduced free 2,4'-bipyridine and 1 '-methyl-

    2,4'-bipyridinium are located at 273 nm and 291 nm respectively. According to Hanasaki

    and Nagakura's work, coordination with a positively charged metal ion will cause a red

    shift of this transition by several thousand wavenumbers. Therefore, the band at 295 nm

    can be assigned as the intraligand Jt(6) to 71(7) transition band. The bands centered at 440

    nm are from metal center to ligand n(7) charge transfer with vibrational coupling.33 The

    band at 268 nm can be assigned as another metal to ligand charge transfer band, from the

    HOMO of platinum(H) ion to the TC(8) orbital of the ligand. The band appearing as

    shoulder at 372 nm with energy about four thousand wavenumbers higher than that of the

    440 nm bands, is also a MLCT band, but it is from the next HOMO's of platinum(II)

    ion center, dz2, to the JI(7) orbital of bipyridinium.36 The origin of the shoulder band at

    343 nm is not very clear; it may result frotti the ligand to metal charger transfer.

    The spectrum of the singly reduced form of this complex strongly suggests that

    this is a ligand-based reduction, the added electron being localized on the bipyridine

    ligand. There is a band at 319 nm in the spectrum of parent species, which is the Jt(6)

    to jc(7) transition band of reduced ligand. This band in the reduced free ligand is at 388

    nm. According to published results,22,29 the blue shift upon coordination is about 5,500

    cm"1, which is consistent with this assignment. The ligand-based it(7) to Jt(10) band,

    which also shifts to higher energy when bipyridines coordinated with platinum(II) ion, is

    located at 473 nm. The broad bands from 538 to 700 nm can not be the MLCT bands,

    because the MLCT bands should move to a higher energy due to the ligands-based

    reduction, but may be assigned as a set of LMCT bands. The band at 375 nm can be

  • 66

    assigned as a MLCT band. The weak bands spreading in the region of 800 to 900 nm

    are from the transition of %(1) to Jt(8), as usual for bipyridine anion radicals. The energy

    and strength of all these bands are closely related to the spectrum of the reduced free

    ligand, so that the first reduction is undoubtedly a ligand-based reduction, as in most

    platinum bipyridine complexes. The band at 260 nm can presumably be assigned as an

    LMCT band, from rc(7) of reduced ligand to the platinum(II) ion center.

    [Pt(bipy)ClJ

    This complex has been known for a long time.1'34 Due to its importance as an

    antitumor agent,21 its properties have been broadly investigated. Gidney reported its UV/-

    VIS spectra in different solvents in 1973.38,39 In 1981, Agarwala assigned its spectrum

    with support from magnetic circular dichroism, and NMR studies.33

    In DMF, the electronic absorption spectra of this compound, shown in Fig.3.23,

    has four bands in the UV/VIS region. As published by Agarwala, the bands at 313 nm

    and 325 nm can be assigned as 2,2'-bipyridine-based it to n* transitions rather than d-d

    or any metal center related charge transfer band. This argument can be supported by their

    extinction coefficient values (e = 7.9 x 103 and 90 x 103) which are far too high to be d-d

    band, and also by the fact that they appear at almost the same position in other mixed-

    ligand platinum(II) 2,2'-bipyridine complexes, such as [Pt(bipy)(4CNpy)Cl]+. The energy

    difference between the two bands is about 1200 cm"1, resulting from vibrational coupling

    of C — C stretching between the two pyridine rings within the same electronic transition.

    The bands at 278 nm and 388 nm are MLCT bands, from platinum(II) to bipyridine-based

    7c(7) and ic(8), respectively. In the spectrum of [Pt(bipy)(4-CNpy)Cl]+, where one of the

  • 67

    4.50

    < 2.25

    1

    •—A-

    800 1000

    Wavelength (nm)

    1200 1400

    Fig.3.23 Electronic absorption spectra of [Pt(bipy)ClJ and its one electron reduction

    product in DMF (c . 2.3 x 10" M) with 0.1 M TBAPF, (Solid line - parent; Dashed

    line -- singly reduced).

  • 68

    chlorides which are 7C donor ligands is replaced by 4-CNpy, a kind of n acceptor ligand,

    these MLCT bands shift to higher energy as the result of lowering the HOMO of the

    metal ion. From the experimental data in Table 3.3, one can see that the energy order

    of Pt(II) —> 7C(7) MLCT bands of Pt(II)BipyL2 complexes is as follow:

    CI < en < MejNpy < py < 4CNpy

    which is roughly consistent with the order of it acceptor ability.

    In the spectrum of this singly reduced species, the sharp band located at 359 nm

    is from the jc(6) —» jr(7) transition of the reduced bipyridine ligand, and the well separated

    three bands at 450, 463 and 498 nm are from the k(1) to Jt(10) transition. The spacing

    between 463 and 498 nm bands is 1500 cm"1, and is of vibrational origin. The band at

    424 nm is not easy to understand; it may come from metal to ligand charge transfer. The

    broad band centered at 913 nm is a typical rc(7) to rc(8) of bipyridine anion radicals.

    Again, the spectrum of this singly reduced complex clearly shows localized bipyridine-

    based reduction.

    [Pt(bipy)(4-CNpy)Cl]+

    The electronic absorption spectra of this compound, as shown in Fig.3.24, is

    simple. There are only two clearly separated bands located at 311 and 323 nm, which

    are the bipyridine-based tc(6) to Jt(7) transition band with vibrational coupling. Due to

    the tc accepting effect of 4-CNpy, the HOMO of platinum is further stabilized so that the

    MLCT band shifts to higher energy, as a shoulder overlapping with the intraligand band

    at 333 nm.

    The spectrum of the singly reduced form of this complex is relatively featureless.

  • 69

    There is a band at 273 nm which possibly is the metal to ligand-based Jt(8) charge

    transfer band. The bands around 320 nm are from the residue of unreduced parent

    compound. There are several bands spreading from 340 to 500 nm, which should include

    MLCT and intraligand n(l) —> tc(10) bands. For some unknown reason, all these bands

    are very weak and further assignment is difficult.

    [PtDCMBClJ

    Fig.3.25 shows the electronic absorption spectra of PtDCMBCl2 in DMF. The

    band at 436 nm originates from metal dxy to ligand rc(7) charger transfer and the shoulder

    band at 373 nm is another MLCT band which is from dz2 or doubly degenerated to

    Jt(7). The separation of the two bands is about 3900 cm'1 which corresponds to the

    energy gap between the two orbitals.

    In the spectrum of the singly reduced form of this complex, there is a Jt(7) to

    7t(8,9) transition, centered at 900 nm, which is weak and broad. The sharp band at 417

    nm is the metal to bipyridine anion radical charge transfer band. The ligand-based k(7)

    to 7t(10) band is at 450 nm.

  • 70

    Wavelength (nm)

    UC absorption spectra of [ P U b i p y K ^ P ^ 1 1 * a n d °"e e l e C t r°"

    reduction product in DMF (c = 1 -2 x

    and singly reduced form (dashed toe)

  • 71

    5.0

    4.5

    4.0

    S 3.0

    J. J.

    600 800 1000

    Wavelength (nm)

    1300

    Fig.3.25 Electronic absorption spectra of [PtDCMBClJ and its one electron reduction

    product in DMF (c = 2.3 x 10"3M) with 0.1 M TBAPF6, parent (solid line) and singly

    reduced (all dashed lines) forms

  • REFERENCES

    (1) Morgan, G. T. ; Burstall, F. H. J. Chem. Soc. 1934, 965.

    (2) (a)Wimmer, S.; Wimmer, F. L. J. Chem. Soc., Dalton Trans. 1994,

    879. (b)Morgan, G. T.; Burstall, F. H. J. Chem. Soc., 1934, 965. (c) Conran,

    R. C.; Rund, J. V. Inorg. Chem., 1972, 11, 129. (d) Dholakia, S.; Gillard, R.

    D.; Wimmer, F. L. Polyhedron, 1985, 4, 791. (e) Howe-Grant, M.; Lippard, S.

    J. Inorg. Synth. 1980, 20, 102.

    (3) Castan, P.; Dahan, F.; Wimmer, S.; Wimmer, F. L. J. Chem. Soc., Dalton

    Trans. 1990, 2971.

    (4) Chassot, L.; Miiller, E.; von Zelewsky, A. Inorg. Chem. 1984, 23, 4249.

    (5) Wimmer, S.; Castan, P.; Wimmer, F. L.; Johnson, N. P. Inorg. Chim. Acta

    1988, 142, 13.

    (6) Matsubayashi, G. E.; Hirao, M.; Tanaka, T. Inorg. Chim. Acta, 1988,

    144, 217.

    (7) Wimmer, S.; Castan, P.; Wimmer, F. L.; Johnson, N. P. J. Chem. Soc., Dalton

    Trans. 1989, 403.

    (8) Sandrini, D.; Maestri, M; Ciano, L.; von Zelewsky, A. Gazz. Chim. Ital.

    1988,118,661.

    (9) Kamath, S. S.; Uma, V.; Srivastava, T. S. Inorg. Chim. Acta 1989, 161, 49.

    (10) Zuleta, J. A.; Burberry, M. S.; Eisenberg, R